We obtain a boundary pointwise gradient estimate on a parabolic half cube Q2R∩{(x1,x′,t)∈Rn+1:x1>0} for nonlinear parabolic equations with measurable nonlinearities, which are only assumed to be measurable in x1-variable. The estimates are obtained in terms of Riesz potential of the right-hand side measure and the oscillation of the boundary data, where the boundary data is given on Q2R∩{(x1,x′,t)∈Rn+1:x1=0}.
Citation: Ho-Sik Lee, Youchan Kim. Boundary Riesz potential estimates for parabolic equations with measurable nonlinearities[J]. Communications in Analysis and Mechanics, 2025, 17(1): 61-99. doi: 10.3934/cam.2025004
[1] | Xiulan Wu, Yaxin Zhao, Xiaoxin Yang . On a singular parabolic $ p $-Laplacian equation with logarithmic nonlinearity. Communications in Analysis and Mechanics, 2024, 16(3): 528-553. doi: 10.3934/cam.2024025 |
[2] | Huiyang Xu . Existence and blow-up of solutions for finitely degenerate semilinear parabolic equations with singular potentials. Communications in Analysis and Mechanics, 2023, 15(2): 132-161. doi: 10.3934/cam.2023008 |
[3] | Yuxuan Chen . Global dynamical behavior of solutions for finite degenerate fourth-order parabolic equations with mean curvature nonlinearity. Communications in Analysis and Mechanics, 2023, 15(4): 658-694. doi: 10.3934/cam.2023033 |
[4] | Farrukh Dekhkonov . On a boundary control problem for a pseudo-parabolic equation. Communications in Analysis and Mechanics, 2023, 15(2): 289-299. doi: 10.3934/cam.2023015 |
[5] | Yue Pang, Xiaotong Qiu, Runzhang Xu, Yanbing Yang . The Cauchy problem for general nonlinear wave equations with doubly dispersive. Communications in Analysis and Mechanics, 2024, 16(2): 416-430. doi: 10.3934/cam.2024019 |
[6] | Farrukh Dekhkonov . On one boundary control problem for a pseudo-parabolic equation in a two-dimensional domain. Communications in Analysis and Mechanics, 2025, 17(1): 1-14. doi: 10.3934/cam.2025001 |
[7] | Yang Liu . Global attractors for a nonlinear plate equation modeling the oscillations of suspension bridges. Communications in Analysis and Mechanics, 2023, 15(3): 436-456. doi: 10.3934/cam.2023021 |
[8] | Yonghui Zou . Global regularity of solutions to the 2D steady compressible Prandtl equations. Communications in Analysis and Mechanics, 2023, 15(4): 695-715. doi: 10.3934/cam.2023034 |
[9] | Panyu Deng, Jun Zheng, Guchuan Zhu . Well-posedness and stability for a nonlinear Euler-Bernoulli beam equation. Communications in Analysis and Mechanics, 2024, 16(1): 193-216. doi: 10.3934/cam.2024009 |
[10] | Sergey A. Rashkovskiy . Nonlinear Pauli equation. Communications in Analysis and Mechanics, 2024, 16(1): 94-120. doi: 10.3934/cam.2024005 |
We obtain a boundary pointwise gradient estimate on a parabolic half cube Q2R∩{(x1,x′,t)∈Rn+1:x1>0} for nonlinear parabolic equations with measurable nonlinearities, which are only assumed to be measurable in x1-variable. The estimates are obtained in terms of Riesz potential of the right-hand side measure and the oscillation of the boundary data, where the boundary data is given on Q2R∩{(x1,x′,t)∈Rn+1:x1=0}.
In this paper, we consider parabolic equations with measurable nonlinearities and measure data
{ut−diva(Du,x1,x′,t)=μinQ+2R,u=ψonT2R, | (1.1) |
where μ is a radon measure with |μ|(Q+2R)<∞. Here, the parabolic half cube with the size ρ is denoted as Q+ρ=(0,ρ)×(−ρ,ρ)n−1×(−ρ2,0) and the parabolic hyperplane with the size ρ is denoted as Tρ={0}×(−ρ,ρ)n−1×(−ρ2,0).
We will obtain the pointwise gradient estimate of u in terms of Riesz potential of the right-hand side μ. Here, Riesz potential of μ is defined as
I|μ|α(x,t,r)=∫r0|μ|(Qρ(x,t))ρn+2−αdρρ((x,t)∈Rn+1,r>0,0<α<n). |
For the boundary data ψ, we measure the pointwise oscillation of the gradient in x′-variable and L2-norm of the time derivative:
∫r0(oscTρDx′ψ+ρ2‖∂tψ‖L2(Tρ))dρρ. |
For the ellipticity constants 0<λ≤Λ, suppose that the nonlinearities a(ξ,x,t) satisfy that
{a(ξ,x,t) is measurable in (x,t) for every ξ∈Rn,a(ξ,x,t) is C1-regular in ξ for almost every (x,t)∈Rn+1, |
and
{|a(ξ,x,t)|≤Λ|ξ|,|Dξa(ξ,x,t)|≤Λ,⟨Dξa(ξ,x,t)ζ,ζ⟩≥λ|ζ|2, | (1.2) |
for any (x,t)∈Rn+1, ξ∈Rn and ζ∈Rn. Also suppose that a(ξ,x,t)=a(ξ,x1,x′,t) is Dini-continuous in (x′,t)-variables, i.e., that there exists a function ω:[0,∞)→[0,1] which is non-decreasing concave with limρ↘0ω(ρ)=ω(0)=0 and
∫10ω(ρ)dρρ<∞, |
satisfying that
|a(ξ,x1,x′,t)−a(ξ,x1,y′,s)|≤ω(|(x′,t)−(y′,s)|)|ξ| | (1.3) |
for every (x1,x′,t)∈Rn+1, ξ∈Rn and y′∈Rn−1. Note that the nonlinearities are only merely measurable on x1-variable.
For nonlinear parabolic equations, many authors obtained the pointwise gradient estimates by using potentials. Duzaar and Mingione considered linear growth condition in [1]. Kuusi and Mingione considered p-growth conditions and obtained Wolff potential type estimates in [2,3] and Riesz potential type estimates in [4]. Also for elliptic equations with measurable nonlinearities, the boundary pointwise gradient estimates by using Riesz potentials were obtained in [5], where they used the excess decay estimates of the gradient in [6]. In this paper, we will extend the result [5] to nonlinear parabolic equations with measurable nonlinearities and obtain the boundary pointwise gradient estimates by using Riesz potentials.
For the reader's further interest, we refer to [7] for Morrey space estimates to linear parabolic systems with measurable coefficients and measure data. We refer to [8] for weighted Lebesgue estimates to linear parabolic systems with measurable coefficients. Potential can be used not only for the right-hand side data of the equation but also be considered as a multiplier of the solution, see for instance [9], which considers existence and blow-up solution with singular potentials multiplied to the solution.
We use the following notations in this paper. Let z be a typical point in Rn, s be a typical time variable and r>0 be a size.
1. z=(z1,⋯,zn)=(z1,z′).
2. Rn+1+={(x1,x′,t)∈Rn+1:x1>0}, Rn+10={(x1,x′,t)∈Rn+1:x1=0}.
3. Qr(z,s)=(z1−r,z1+r)×⋯×(zn−r,zn+r)×(s−r2,s), Qr=Qr(0,0).
4. Q+r(z,s)=Qr(z,s)∩Rn+1+, Q+r=Qr∩Rn+1+.
5. Tr(z′,s)={0}×(z2−r,z2+r)×⋯×(zn−r,zn+r)×(s−r2,s)=Qr(0,z′,s)∩Rn+10.
6. Kr(z)=(z1−r,z1+r)×⋯×(zn−r,zn+r).
7. K+r(z)=(z1−r,z1+r)×⋯×(zn−r,zn+r)∩Rn+.
8. For g∈L1(U), (g)U=−∫Ugdx=1|U|∫Ugdx when |U|≠0.
In view of the available approximation theory, we assume that μ∈L1(Q+2R). Without loss of generality, we shall assume that
μ∈L1(Rn+1), | (1.4) |
by letting μ|Rn+1∖Q+2R=0. By using the concept of SOLA (Solutions Obtained by Limit of Approximations), we will remove this assumption in Corollary 1.3. Also for the boundary data, let ψ:T2R→R be a function such that
Dx′ψ∈L∞(T2R)and∂tψ∈L2(T2R). | (1.5) |
We obtain the following boundary pointwise gradient estimate in this paper.
Theorem 1.1. There exists a constant c1=c1(n,λ,Λ)≥1 such that the following holds. For some r∈(0,R], assume that
c1∫2r0ω(ρ)dρρ≤1. | (1.6) |
If u∈C0(−4R2,0;L2(K+2R))∩L2(−4R2,0;W1,1(K+2R)) is a weak solution of
{ut−diva(Du,x1,x′,t)=μinQ+2R,u=ψonT2R | (1.7) |
with the assumptions (1.2)–(1.5), then we have that
|Du(x0,t0)|≤c[−∫Q+r(x0,t0)|Du|dxdt+|Dx′ψ(x′0,t0)|]+c[∫2r0(|μ|(Q+ρ(x0,t0))ρn+1+oscTρ(x′0,t0)Dx′ψ+ρ2−∫Q+ρ(x0,t0)|∂tψ|2dxdt)dρρ], | (1.8) |
for any Lebesgue point (x0,t0)=(x10,x′0,t0)∈¯Q+R of Du with c=c(n,λ,Λ).
To deal with SOLA, we now remove the assumption (1.4).
Definition 1.2. A SOLA of (1.7) is a distributional solution u∈L2(−4R2,0;W1,1(K+2R)) to (1.7) such that u is the limit of solutions uh∈C0(−4R2,0;L2(K+2R))∩L2(−4R2,0;W1,1(K+2R)) to
{(uh)t−diva(Duh,x1,x′,t)=μh in Q+2R,uh=ψ on T2R, |
in the sense that uh→u in L2(−4R2,0;W1,1(K+2R)) and L∞∋μh⇀μ in the sense of measures satisfying
lim suph→∞|μh|(Q+ρ(x0,t0))≤|μ|(⌊Q+ρ(x0,t0)⌋) for any Q+ρ(x0,t0)⊂Q+2R, |
where ⌊Q⌋ denotes the parabolic closure of Q.
We finally state our main result for SOLA.
Corollary 1.3. Without (1.4), Theorem 1.1 continues to hold for SOLA of (1.7), with the estimate (1.8) for any Lebesgue point (x0,t0)=(x10,x′0,t0)∈¯Q+R of Du.
Remark 1.4. For the sake of convenience and simplicity, we employ the letter c>0 and α∈(0,1] throughout this paper to denote any constants which can be explicitly computed in terms of known quantities such as n,λ,Λ. Thus the exact values denoted by c and α may change from line to line in each given computation.
In this section, the nonlinearities are assumed to be depending only on ξ and x1-variables with the following assumptions:
{a(ξ,x1) is measurable in x1∈R for every ξ∈Rn,a(ξ,x1) is C1 −regular in ξ∈Rn for every x1∈R. | (2.1) |
Also we assume that a(ξ,x1):Rn×R→Rn satisfies
{|a(ξ,x1)|≤Λ|ξ|+Γ,|Dξa(ξ,x1)|≤Λ,⟨Dξa(ξ,x1)ζ,ζ⟩≥λ|ζ|2, | (2.2) |
for every x1∈R, ξ∈Rn, ζ∈Rn and for some constants 0<λ≤Λ, Γ≥0.
We obtain the boundary excess-decay estimates for parabolic equations in this section. Under the assumptions (2.1), (2.2) and
0≤x10≤r, | (2.3) |
let g be a weak solution of
{gt−div a(Dg,x1)=0 in Q+3r(x0,t0),g=γ′⋅x′ on T3r(x′0,t0), | (2.4) |
where γ′=(γ2,⋯,γn)∈Rn−1 and (x0,t0)=(x10,x20,⋯,xn0)=(x10,x′0,t0).
Remark 2.1. If the nonlinearity a(ξ,x1) is only defined on 0<x1<3r, then one can easily extend a(ξ,x1) to satisfy (2.1) and (2.2), which does not effect the results in this paper.
The nonlinearity a(ξ,x1) only depends on x1-variable and g=γ′⋅x′ on T3r(x′0,t0). So one can use the difference quotient method to find that Dkg−γk (k∈{2,3,⋯,n}) is weakly differentiable in Q+2r(x0,t0). Moreover, one can show that Dkg−γk∈W1,2(Q+2r(x0,t0)). By differentiating (2.4) with respect to xk-variable, one can get that
{∂t(Dkg−γk)−Di[aij(x,t)Dj(Dkg−γk)]=0 in Q+2r(x0,t0),Dkg−γk=0 on T2r(x′0,t0), | (2.5) |
where aij(x,t)=∂ai∂ξj(Dg,x1) satisfies that
{aij(x,t)ζiζj≥λ|ζ|2,|aij(x,t)|≤Λ, | (2.6) |
for any (x,t)∈Q+2r(x0,t0) and ζ∈Rn with 1≤i,j≤n.
To obtain boundary estimates, we next extend the equation (2.5) from Q+2r(x0,t0) to Q2r(x0,t0). For k∈{2,3,⋯,n}, we let
gk be the odd extension of Dkg−γk from Q+2r(x0,t0) to Q2r(x0,t0). | (2.7) |
Then
{∂tgk−Di[aij(x,t)Djgk]=0 in Q+2r(x0,t0),gk=0 on T2r(x′0,t0). | (2.8) |
Let ˆaij(x,t) be an extension of aij(x,t) from Q+2r(x0,t0) to Q2r(x0,t0) defined as
{ˆa11(−x1,x′,t)=a11(x1,x′,t),ˆaij(−x1,x′,t)=aij(x1,x′,t)when 1<i≤n,1<j≤n,ˆa1j(−x1,x′,t)=−a1j(x1,x′,t)when 1<j≤n,ˆai1(−x1,x′,t)=−ai1(x1,x′,t)when 1<i≤n, | (2.9) |
for (x1,x′,t)∈Q2r(x0,t0)∖Q+2r(x0,t0). Then one can check from (2.6) that
{ˆaij(x,t)ζiζj≥λ|ζ|2,|ˆaij(x,t)|≤Λ, | (2.10) |
for any (x,t)∈Q2r(x0,t0) and ζ∈Rn. Then we obtain from (2.8) and (2.9) that gk is a weak solution of the parabolic equation
∂tgk−Di[ˆaij(x,t)Djgk]=0 in Q2r(x0,t0). | (2.11) |
From [10, Chapter 6], we have an excess decay estimate for linear parabolic equations, which can be applied to (2.11).
Lemma 2.2. Under the assumptions
{aij(x,t)ζiζj≥λ|ζ|2((x,t)∈Qr,ζ∈Rn),‖aij‖L∞(Qr)≤Λ, |
let w be a weak solution of
∂tw−Di[aij(x,t)Djw]=0inQr. |
Then we have that
−∫Qρ|w−(w)Qρ|2dxdt≤c(ρr)2α−∫Qr|w−(w)Qr|2dxdt(0<ρ≤r), |
where c=c(n,λ,Λ) and α=α(n,λ,Λ)∈(0,1].
With Lemma 2.2 and the energy estimate, we obtain the following lemma.
Lemma 2.3. Suppose that k∈{2,3,⋯,n}. For gk in (2.7), we have that
−∫Qτ(y,s)|gk−(gk)Qτ(y,s)|2dxdt≤c(τρ)2α−∫Qρ(y,s)|gk−(gk)Qρ(y,s)|2dxdt, | (2.12) |
and
∫Qτ(y,s)|Dgk|2dxdt≤c(ρ−τ)2∫Qρ(y,s)|gk−ζk|2dxdt(ζk∈R), | (2.13) |
for any Qρ(y,s)⊂Q2r(x0,t0) and 0<τ<ρ.
Proof. Let k∈{2,3,⋯,n} be an arbitrary integer. The estimate (2.12) follows by applying Lemma 2.2 and (2.10) to (2.11).
Next, we choose a cut-off function η∈C∞c(Qρ(y,s)) with
0≤η≤1,η=1 on Qτ(y,s),|Dη|≤cρ−τ and |∂tη|≤c(ρ−τ)2. | (2.14) |
We have from (2.11) that
∂t(gk−ζk)−Di[ˆaij(x,t)Dj(gk−ζk)]=0 in Q2r(x0,t0). | (2.15) |
Since η∈C∞c(Qρ(y,s)) and Qρ(y,s)⊂Q2r(x0,t0), we test the above equation by [gk−ζk]η2 to find that
0=∫Qρ(y,s)[∂t{([gk−ζk]η)22}−[gk−ζk]2η∂tη]dxdt+∫Qρ(y,s)ˆaij(x,t)DjgkDi{[gk−ζk]η2}dxdt. |
Since η∈C∞c(Qρ(y,s)), one can check that ∫Qρ(y,s)∂t{([gk−ζk]η)2}dxdt≥0. So by (2.10),
λ∫Qρ(y,s)|Dgk|2η2dxdt≤∫Qρ(y,s)[∂t{([gk−ζk]η)22}+ˆaij(x,t)DjgkDigkη2]dxdt=∫Qρ(y,s){[gk−ζk]2η∂tη−ˆaij(x,t)Djgk[gk−ζk]2ηDiη}dxdt≤c∫Qρ(y,s){|gk−ζk|2|η||∂tη|+|Dgk||gk−ζk||η||Dη|}dxdt. |
By Young's inequality,
∫Qρ(y,s)|Dgk|2η2dxdt≤c∫Qρ(y,s)|gk−ζk|2{|Dη|2+|η||∂tη|}dxdt. |
So (2.13) follows from (2.14).
Now, we extend gt from Q+2r(x0,t0) to Q2r(x0,t0) and obtain some estimates on the extended function of gt. Since g is a weak solution of (2.4), one can use the difference quotient method to show that gt=∂tg is a weak solution of
(gt)t−Di[aij(x,t)Dj(gt)]=∂t[gt−Di{ai(Dg,x1)}]=0 in Q+2r(x0,t0). | (2.16) |
In view of (2.4), one can check that gt=0 on T2r(x′0,t0). So
let gn+1 be the odd extension of gt from Q+2r(x0,t0) to Q2r(x0,t0) | (2.17) |
defined as
gn+1(x1,x′)={gt(x1,x′) in Q+2r(x0,t0),−gt(−x1,x′) in Q2r(x0,t0)∖Q+2r(x0,t0). |
So we find from (2.9) and (2.16) that
∂tgn+1−Di[ˆaij(x,t)Djgn+1]=0 in Q2r(x0,t0). | (2.18) |
Then we have the following energy estimate for gn+1 in (2.18).
Lemma 2.4. If gn+1 is a weak solution of (2.18), then we have that
∫Qτ(y,s)|Dgn+1|2dxdt≤c(ρ−τ)2∫Qρ(y,s)|gn+1|2dxdt(0<τ<ρ), |
for any Qρ(y,s)⊂Q2r(x0,t0).
Proof. Choose a cut-off function η∈C∞c(Qρ(y,s)) with
0≤η≤1,η=1 in Qτ(y,s),|Dη|≤cρ−τand|∂tη|≤c(ρ−τ)2. | (2.19) |
We test (2.18) by φ=η2gn+1 to find that
0=∫Qρ(y,s)[∂t(gn+1)gn+1η2+ˆaij(x,t)Djgn+1Di(η2gn+1)]dxdt. | (2.20) |
Then a direct calculation gives that
λ∫Qρ(y,s)|Dgn+1|2η2dxdt≤∫Qρ(y,s)ˆaij(x,t)Dj(gn+1)η2Di(gn+1)dxdt=∫Qρ(y,s)ˆaij(x,t)Dj(gn+1)[Di(η2gn+1)−2ηDiηgn+1]dxdt=−∫Qρ(y,s)[∂t(gn+1)η2gn+1+ˆaij(x,t)Dj(gn+1)2ηDiηgn+1]dxdt. |
From the fact that η∈C∞c(Qρ(y,s)), we get
∫Qρ(y,s)∂t(gn+1)η2gn+1dxdt=∫Qρ(y,s)∂t(g2n+1η2)2−|gn+1|2ηηtdxdt≥−∫Qρ(y,s)|gn+1|2ηηtdxdt. |
By combining the above two equalities and applying the elliptic condition (2.6), we get
∫Qρ(y,s)|Dgn+1|2η2dxdt≤c∫Qρ(y,s)|gn+1|2(|Dη|2+|ηηt|)dxdt. |
So the lemma follows from the choice of the cut-off function η in (2.19).
Since the nonlinearity a(ξ,x1) depends on x1-variable, we obtain an excess decay estimate in terms of a1(Dg,x1) instead of D1g. Let g1 be the even extension of a1(Dg,x1) from Q+2r(x0,t0) to Q2r(x0,t0) defined as
{g1(x1,x′,t)=a1(Dg(x1,x′,t),x1) in Q+2r(x0,t0)g1(x1,x′,t)=a1(Dg(−x1,x′,t),−x1) in Q2r(x0,t0)∖Q+2r(x0,t0). | (2.21) |
So from (2.7), (2.17) and (2.21), we have following extensions from Q+2r(x0,t0) to Q2r(x0,t0):
{g1 is the even extension of a1(Dg,x1),gk (k∈{2,3,⋯,n}) is the odd extension of Dkg−γk,gn+1 is the odd extension of gt. | (2.22) |
Then we define G:Q2r(x0,t0)→Rn as
G=(g1,g2,⋯,gn). | (2.23) |
The desired excess-decay estimate will be obtained with the function G in (2.23).
With Lemma 2.4, we estimate gn+1 by using the function G in (2.23).
Lemma 2.5. For gn+1 in (2.17), we have that
∫Qτ(y,s)|gn+1|2dxdt≤c(ρ−τ)2∫Qρ(y,s)|G−ζ|2dxdt(0<τ<ρ), |
for any Qρ(y,s)⊂Q2r(x0,t0) and ζ=(ζ1,ζ2,⋯,ζn)∈Rn.
Proof. Let d0=τ and d∞=ρ. Let
dm=d0+m∑l=1ρ−τ2landem−1=dm−1+dm2(m=1,2,3,⋯). |
Choose a cut-off function η∈C∞c(Qem(x0,t0)) with
0≤η≤1,η=1 in Qdm(x0,t0),|Dη|≤c2mr−ρand|∂tη|≤c4m(r−ρ)2. | (2.24) |
Since gt=0 on T2r(x′0,t0), we test (2.4) by η2gt to find that
∫Q+em(y,s)gt[η2gt]dxdt=−∫Q+em(y,s)⟨a(Dg,x1),D[η2gt]⟩dxdt. | (2.25) |
Since η∈C∞c(Qem(x0,t0)) and gt=0 on T2r(x′0,t0), one can easily check that
∫Q+em(y,s)a1(Dg,x1)D1[η2gt]dxdt=∫Q+em(y,s)[a1(Dg,x1)−ζ1]D1[η2gt]dxdt. |
Then for κ>0, Young's inequality implies that
|∫Q+em(y,s)a1(Dg,x1)D1[η2gt]dxdt|≤∫Q+em(y,s)[κ|Dgt|2η2+η2|gt|248+c(η2κ+|Dη|2)|a1(Dg,x1)−ζ1|2]dxdt. | (2.26) |
By using integration by parts, for any k∈{2,3,⋯,n} we have that
∫Q+em(y,s)ak(Dg,x1)Dk[η2gt]dxdt=−∫Q+em(y,s)Dk[ak(Dg,x1)]η2gtdxdt. |
Then Young's inequality implies that
n∑k=2|∫Q+em(y,s)ak(Dg,x1)Dk[η2gt]dxdt|≤∫Q+em(y,s)[η2|gt|248+cn∑k=2|Dk[ak(Dg,x1)]|2η2]dxdt. |
By (2.7), gk is the odd extension of Dkg−γk from Q+2r(x0,t0) to Q2r(x0,t0), which implies
n∑k=2∫Q+em(y,s)|Dk[ak(Dg,x1)]|2η2dxdt≤cn∑k=2∫Q+em(y,s)|DDkg|2η2dxdt≤cn∑k=2∫Q+em(y,s)|Dgk|2η2dxdt. |
By combining the above two estimates, we apply τ=em and ρ=dm+1 in Lemma 2.3. Then
|n∑k=2∫Q+em(y,s)ak(Dg,x1)Dk[η2gt]dxdt|≤∫Q+em(y,s)η2|gt|248dxdt+∫Q+dm+1(y,s)c4m(ρ−τ)2n∑k=2|gk−ζk|2dxdt, | (2.27) |
because dm+1−em+1=dm+1−dm2=ρ−τ2m+1. So we obtain from (2.25), (2.26) and (2.27) that
∫Q+em(y,s)η2|gt|2dxdt≤∫Q+em(y,s)[κ|Dgt|2η2+η2|gt|224]dxdt+c∫Q+dm+1(y,s)[(η2κ+|Dη|2)|a1(Dg,x1)−ζ1|2+4m(ρ−τ)2n∑k=2|gk−ζk|2]dxdt. |
By (2.22), g1 is the even extension of a1(Dg,x1), gk (k∈{2,3,⋯,n}) is the odd extension of Dkg−γk and gn+1 is the odd extension of gt from Q+2r(x0,t0) to Q2r(x0,t0). Thus
∫Qem(y,s)η2|gn+1|2dxdt≤∫Qem(y,s)[2κ|Dgn+1|2η2+η2|gn+1|212]dxdt+c∫Qdm+1(y,s)[(η2κ+|Dη|2)|g1−ζ1|2+4m(ρ−τ)2n∑k=2|gk−ζk|2]dxdt. | (2.28) |
Now, take ρ=em and r=dm+1 in Lemma 2.4 to find that
∫Qem(y,s)|Dgn+1|2dxdt≤c14m(ρ−τ)2∫Qdm+1(y,s)|gn+1|2dxdt. | (2.29) |
Take κ so that c1κ4m(r−ρ)2=148. By combining (2.24), (2.28) and (2.29), we have
∫Qdm(y,s)|gn+1|2dxdt≤∫Qdm+1(y,s)|gn+1|28+4mc|G−ζ|2(ρ−τ)2dxdt. | (2.30) |
Multiply (2.30) by 18m and sum it from m=0 to ∞. Then we have that
∞∑m=018m∫Qdm(y,s)|gn+1|2dxdt≤∞∑m=018m+1∫Qdm+1(y,s)|gn+1|2dxdt+∞∑m=0c2m(ρ−τ)2∫Qρ(y,s)|G−ζ|2dxdt. | (2.31) |
Thus from (2.31) and the fact that d0=τ, we have
∫Qτ(y,s)|gn+1|2dxdt=∫Qd0(y,s)|gn+1|2dxdt≤c(ρ−τ)2∫Qρ(y,s)|G−ζ|2dxdt, |
which finishes the proof.
To obtain the desired excess-decay estimate on G=(g1,⋯,gn), we will use Poincaré's inequality. In Lemma 2.3 and Lemma 2.4, the derivatives Dg2,⋯,Dgn and Dgn+1 were obtained. So it only remains to obtain the following estimate on Dg1.
Lemma 2.6. For g1 in (2.22), Dg1∈L2(Qr(x0,t0)) exists with the estimate
|Dg1|≤c(∑2≤k≤n|Dgk|+|gn+1|)∈L2(Qr(x0,t0)). |
Proof. We discover from Lemma 2.3 and Lemma 2.5 that
Dgk∈L2(Qr(x0,t0)) and gn+1∈L2(Qr(x0,t0)), |
for any k∈{2,3,⋯,n}. It follows from (2.22) that
DDx′g∈L2(Q+r(x0,t0)) and gt∈L2(Q+r(x0,t0)). | (2.32) |
Since g is a weak solution of (2.4) and the nonlinearities a(ξ,x1) are independent of xk-variable for any k∈{2,3,⋯,n}, we have from (2.32) that
{D1[a1(Dg,x1)]=gt−n∑k=2Dk[ak(Dg,x1)]=gt−n∑k=2akj(x,t)Djkg∈L2(Q+r(x0,t0))Dx′[a1(Dg,x1)]=a1j(x,t)DjDx′g∈L2(Q+r(x0,t0)). |
From (2.6), aij(x,t) is uniformly elliptic. So we find from (2.22) that
|D[a1(Dg,x1)]|≤c(|DDx′g|+|gt|)≤c(∑2≤k≤n|Dgk|+|gn+1|)∈L2(Q+r(x0,t0)). |
By (2.22), g1 is the even extension of a1(Dg,x1), gk (k∈{2,3,⋯,n}) is the odd extension of Dkg−γk and gn+1 is the odd extension of gt from Q+2r(x0,t0) to Q2r(x0,t0). So the lemma follows by extending the above estimate from Q+r(x0,t0) to Qr(x0,t0).
We obtain the following excess-decay estimate and L∞-estimate of gn+1.
Lemma 2.7. For the odd extension gn+1 of gt in (2.22), we have that
−∫Qτ(y,s)|gn+1−(gn+1)Qτ(y,s)|2dxdt≤c(ρr)2α−∫Qρ(y,s)|gn+1−(gn+1)Qρ(y,s)|2dxdt, |
and
‖gn+1‖2L∞(Qρ2(y,s))≤c−∫Qρ(y,s)|gn+1|2dxdt. | (2.33) |
for any Qρ(y,s)⊂Qr(x0,t0) and 0<τ≤ρ where α=α(n,λ,Λ)∈(0,1].
Proof. From (2.18), gn+1 is a weak solution of
∂tgn+1−Di[aij(x,t)Djgn+1]=0 in Qr(x0,t0). | (2.34) |
By using (2.6) and applying Lemma 2.2 to (2.34), we find that
−∫Qτ(y,s)|gn+1−(gn+1)Qτ(y,s)|2dxdt≤c(ρr)2α−∫Qρ(y,s)|gn+1−(gn+1)Qρ(y,s)|2dxdt, | (2.35) |
for any Qρ(y,s)⊂Qr(x0,t0) and 0<τ≤ρ. The L∞-estimate of gn+1 in (2.33) follows by applying Campanato type embedding to the excess-decay estimate (2.35).
Recall the definition of G in (2.22) and (2.23). In view of Lemma 2.4 and Lemma 2.6, one can estimate DG and ∂tG as
{|DG|≤c{∑1≤k≤n|Dgk|}≤c{∑2≤k≤n|Dgk|+|gn+1|},|∂tG|≤c|Dgn+1|, | (2.36) |
which implies that
−∫Qρ(y,s){|DG|2+ρ2|∂tG|2}dxdt≤c−∫Qρ(y,s){∑2≤k≤n|Dgk|2+|gn+1|2+ρ2|Dgn+1|2}dxdt, | (2.37) |
for any Qρ(y,s)⊂Qr(x0,t0). Here, Dgk (k∈{2,3,⋯,n}), gn+1 and Dgn+1 were estimated in Lemma 2.3, Lemma 2.5 and Lemma 2.4 respectively. So we use Sobolev type embeddings to have the following reverse Hölder type inequality.
Lemma 2.8. For G in (2.22) and (2.23), we have that
(−∫Qρ2(y,s)|G−ζ|2∗dxdt)12∗≤c−∫Qρ(y,s)|G−ζ|dxdt(ζ∈Rn), |
for any Qρ(y,s)⊂Qr(x0,t0). Here, 2∗=2(n+1)n−1>2 is the Sobolev conjugate for (n+1)-dimension.
Proof. Fix any Qρ(y,s)⊂Qr(x0,t0). Choose arbitrary constants ρ2≤τ1<τ2≤ρ. Then by the Sobolev type embedding,
(−∫Qτ1(y,s)|G−ζ|2∗dxdt)22∗≤c−∫Qτ1(y,s){τ21|DG|2+τ41|∂tG|2+|G−ζ|2}dxdt. |
Here, the Sobolev conjugate for (n+1)-dimension is denoted as 2∗=2(n+1)n−1>2. We have from (2.37) that
−∫Qτ1(y,s)[τ21|DG|2+τ41|∂tG|2]dxdt≤cτ21−∫Qτ1(y,s)[∑2≤k≤n|Dgk|2+|gn+1|2+τ21|Dgn+1|2]dxdt. |
Since ρ2≤τ1<τ2≤ρ, we have from (2.13) in Lemma 2.3 and Lemma 2.5 that
−∫Qτ1(y,s)[∑2≤k≤n|Dgk|2+|gn+1|2]dxdt≤c(τ2−τ1)2−∫Qτ2(y,s)[∑2≤k≤n|gk−ζk|2+|G−ζ|2]dxdt≤c(τ2−τ1)2−∫Qτ2(y,s)|G−ζ|2dxdt. |
Since ρ2≤τ1<τ2≤ρ, we have from Lemma 2.4 and Lemma 2.5 that
−∫Qτ1(y,s)|Dgn+1|2dxdt≤c(τ2−τ1)2−∫Qτ1+τ22(y,s)|gn+1|2dxdt≤c(τ2−τ1)4−∫Qτ2(y,s)|G−ζ|2dxdt. |
Since τ21(τ2−τ1)2≥1, by combining the above four estimates, we get
(−∫Qτ1(y,s)|G−ζ|2∗dxdt)22∗≤cτ41(τ2−τ1)4−∫Qτ2(y,s)|G−ζ|2dxdt. |
By the interpolation inequality, we get
−∫Qτ2(y,s)|G−ζ|2dxdt≤(−∫Qτ2(y,s)|G−ζ|2(n+1)n−1dxdt)n−1n+3(−∫Qτ2(y,s)|G−ζ|dxdt)4n+3. |
Since 2∗=2(n+1)n−1>2, we obtain from Young's inequality that
(−∫Qτ1(y,s)|G−ζ|2∗dxdt)22∗≤12(−∫Qτ2(y,s)|G−ζ|2∗dxdt)22∗+cτ2(n+3)1(τ2−τ1)2(n+3)(−∫Qτ2(y,s)|G−ζ|dxdt)2. |
Since ρ2≤τ1<τ2≤ρ were chosen arbitrarily, by [11, Lemma 4.3], we get
(−∫Qρ2(y,s)|G−ζ|2∗dxdt)22∗≤c(−∫Qρ(y,s)|G−ζ|dxdt)2, |
and the lemma follows.
By using (2.37), we apply Poincaré's inequality to obtain the desired excess-decay estimate on G=(g1,⋯,gn). We remark that Dg1, Dgk (k∈{2,⋯,n}), Dgn+1 and gn+1 were estimated in Lemma 2.6, Lemma 2.3, Lemma 2.4 and Lemma 2.5 respectively.
Lemma 2.9. For G in (2.22) and (2.23), we have that
−∫Qτ(y,s)|G−(G)Qτ(y,s)|dxdt≤c(τρ)α−∫Qρ(y,s)|G−(G)Qρ(y,s)|dxdt(0<τ≤ρ), |
for any Qρ(y,s)⊂Qr(x0,t0) where α=α(n,λ,Λ)∈(0,1].
Proof. Assume that 8τ≤ρ, otherwise the lemma holds from the dilation. We claim that
−∫Qτ(y,s)|G−(G)Qτ(y,s)|2dxdt≤c(τρ)2α−∫Qρ2(y,s)|G−(G)Qρ(y,s)|2dxdt. | (2.38) |
From Poincaré's inequality and (2.37), we have that
−∫Qτ(y,s)|G−(G)Qτ(y,s)|2dxdt≤c−∫Qτ(y,s){τ2|DG|2+τ4|∂tG|2}dxdt.≤cτ2−∫Qτ(y,s){∑2≤k≤n|Dgk|2+|gn+1|2+τ2|Dgn+1|2}dxdt. | (2.39) |
By taking ζk=(gk)Q2τ(y,s) in Lemma 2.3, we get
τ2−∫Qτ(y,s){∑2≤k≤n|Dgk|2+τ2|gn+1|2}dxdt≤c−∫Q2τ(y,s){∑2≤k≤n|gk−(gk)Q2τ(y,s)|2+τ4‖gn+1‖2L∞(Qτ(y,s))}dxdt. |
By the assumption 8τ≤ρ, we have from (2.12) in Lemma 2.3 that
∑2≤k≤n−∫Q2τ(y,s)|gk−(gk)Q2τ(y,s)|2dxdt≤c(τρ)2α∑2≤k≤n−∫Qρ2(y,s)|gk−(gk)Qρ2(y,s)|2dxdt≤c(τρ)2α−∫Qρ2(y,s)|G−(G)Qρ2(y,s)|2dxdt≤c(τρ)2α−∫Qρ2(y,s)|G−(G)Qρ(y,s)|2dxdt. |
By combining the above two estimates, we get
τ2−∫Qτ(y,s){∑2≤k≤n|Dgk|2+τ2|gn+1|2}dxdt≤c[(τρ)2α−∫Qρ2(y,s)|G−(G)Qρ(y,s)|2dxdt+τ4‖gn+1‖2L∞(Qτ(y,s))]. | (2.40) |
By the assumption 8τ≤ρ, we have from (2.33) in Lemma 2.7 that
τ4‖gn+1‖2L∞(Qτ(y,s))≤τ4‖gn+1‖2L∞(Qρ8(y,s))≤cτ4−∫Qρ4(y,s)|gn+1|2dxdt. |
Also we take ζ=(G)Qρ2(y,s) in Lemma 2.5 to find that
τ4−∫Qρ4(y,s)|gn+1|2dxdt≤cτ4ρ2−∫Qρ2(y,s)|G−(G)Qρ2(y,s)|2dxdt≤cτ4ρ2−∫Qρ2(y,s)|G−(G)Qρ(y,s)|2dxdt. |
By combining the above two estimates, we get
τ4‖gn+1‖2L∞(Qτ(y,s))≤cτ4ρ2−∫Qρ2(y,s)|G−(G)Qρ(y,s)|2dxdt. | (2.41) |
The claim (2.38) holds from (2.39), (2.40) and (2.41). With Hölder's inequality, the lemma follows from (2.38) and Lemma 2.8 by taking ζ=(G)Qρ(y,s).
For the comparison estimates on Qr(x0,t0), we handle the interior case x10>r in Subsection 3.1 and the boundary case 0≤x10≤r in Subsection 3.2. From [12, Lemma 4.1], the absolute value of measurable nonlinearities |a(ξ,x1)| is comparable to |ξ|. In fact, one can easily modify the proof of [12, Lemma 4.1] to obtain the following result, where the nonlinearities depend on ξ, x and t.
Lemma 3.1. Suppose that (1.2). For any (x,t)∈Rn+1, we have that
|ξ|≤c[|ξ′|+(2Λ)−1|a1(ξ,x,t)|]≤c|ξ|(ξ=(ξ1,ξ′)∈Rn). |
For a weak solution u of
ut−div a(Du,x1,x′,t)=μ in Qr(x0,t0), |
let v and g be the weak solution of
{vt−div a(Dv,x1,x′,t)=0 in Qr(x0,t0),v=u on ∂pQr(x0,t0), |
and
{gt−div a(Dg,x1,x′0,t0)=0 in Qr2(x0,t0),g=v on ∂pQr2(x0,t0), |
where ∂p denotes the parabolic boundary. By repeating the proof of the comparison estimate for Du and Dv such as in [1, Lemma 4.1] and [3, Lemma 4.1], one can prove that
−∫Qr(x0,t0)|Du−Dv|dx≤c|μ|(Qr(x0,t0))rn+1. |
By repeating the proof such as in [1, Lemma 4.2], one can prove that
−∫Qr2(x0,t0)|Dv−Dg|dx≤cω(r)−∫Qr(x0,t0)|Dv|dx. |
So we obtain that
−∫Qr2(x0,t0)|Du−Dg|dx≤c[|μ|(Qr(x0,t0))rn+1+ω(r)−∫Qr(x0,t0)|Du|dx]. | (3.1) |
We set
{U=(a1(Du,x1,x′0,t0),D2u,⋯,Dnu),G=(a1(Dg,x1,x′0,t0),D2g,⋯,Dng). | (3.2) |
From [13, Lemma 4.9], we have that
−∫Qρ(x0,t0)|G−(G)Qρ(x0,t0)|dx≤c(ρr)α−∫Qr2(x0,t0)|G−(G)Qr2(x0,t0)|dx(0<2ρ≤r). | (3.3) |
From Lemma 3.1, we have that |Du|≤c|U|. Since |U−G|≤c|Du−Dg|, we find from (3.1) that
−∫Qr2(x0,t0)|U−G|dx≤c[|μ|(Qr(x0,t0))rn+1+ω(r)−∫Qr(x0,t0)|Du|dx]≤c[|μ|(Qr(x0,t0))rn+1+ω(r)−∫Qr(x0,t0)|U|dx]. |
So we obtain from (3.3) that
−∫Qρ(x0,t0)|U−(U)Qρ(x0,t0)|dx≤c(ρr)α−∫Qr2(x0,t0)|U−(U)Qr2(x0,t0)|dx+c(rρ)n[|μ|(Qr(x0,t0))rn+1+ω(r)−∫Qr(x0,t0)|U|dx]. |
One can easily check that
−∫Qr2(x0,t0)|U−(U)Qr2(x0,t0)|dx≤−∫Qr2(x0,t0)|U−(U)Qr(x0,t0)|dx+−∫Qr2(x0,t0)|(U)Qr(x0,t0)−(U)Qr2(x0,t0)|dx≤2n+1−∫Qr(x0,t0)|U−(U)Qr(x0,t0)|dx |
so that
−∫Qρ(x0,t0)|U−(U)Qρ(x0,t0)|dx≤c(ρr)α−∫Qr(x0,t0)|U−(U)Qr(x0,t0)|dx+c(rρ)n[|μ|(Qr(x0,t0))rn+1+ω(r)−∫Qr(x0,t0)|U|dx] | (3.4) |
for any 0<2ρ≤r.
To handle the boundary case, we assume that
0≤x10≤r. | (3.5) |
For the boundary data, we assume that
Dx′ψ∈L∞(T4r(x′0,t0))and∂tψ∈L2(T4r(x′0,t0)). |
We regard the boundary data ψ as a function in Q+4r(x0,t0) by defining ψ(x1,x′,t)=ψ(0,x′,t) for every 0<x1<4r. For a weak solution u of
{ut−div a(Du,x1,x′,t)=μ in Q+4r(x0,t0),u=ψ on T4r(x′0,t0), | (3.6) |
let v, w and g be the weak solution of
{vt−div a(Dv,x1,x′,t)=0 in Q+4r(x0,t0),v=u on ∂p[Q+4r(x0,t0)], | (3.7) |
{wt−div a(Dw,x1,x′,t)=0 in Q+4r(x0,t0),w=v−ψ+Dx′ψ(x′0,t0)⋅x′ on ∂p[Q+4r(x0,t0)], | (3.8) |
and
{gt−div a(Dg,x1,x′0,t0)=0 in Q+3r(x0,t0),g=w on ∂p[Q+3r(x0,t0)]. | (3.9) |
We have from (3.6) and (3.7) that v=u=ψ on T4r(x′0,t0). So from (3.7) and (3.8), we have that
w=Dx′ψ(x′0,t0)⋅x′ on T4r(x′0,t0). | (3.10) |
By following the proof of [3, Lemma 4.1] (although [3] considered p-Laplace type equations), we obtain the comparison estimate for Du and Dv to our problems. The proof for Lemma 3.2 is similar to that of [3, Lemma 4.1], but we give the proof for the convenience of the readers.
Lemma 3.2. Under the assumption (3.5), we have that
−∫Q+4r(x0,t0)|Du−Dv|dx≤c|μ|(Q+4r(x0,t0))rn+1. |
Proof. We first claim that
supτ∈(t0−r2,t0)∫K4r(x0)|u(x,τ)−v(x,τ)|dx≤|μ|(Q+4r(x0,t0)). | (3.11) |
To prove the claim (3.11), fix τ∈(t0−r2,t0). For m≥1, let ϕ:R→[0,1] be a function only depending on t-variable defined as
ϕ(t)={0 if t≥τ,m(τ−t) if τ−1m≤t<τ,1 if t<τ−1m. | (3.12) |
Here, we remark that we will let m→∞ later. We also define
η1,ϵ=±min{1,(u−v)±ϵ}ϕ(ϵ>0), |
which implies that
Dη1,ϵ=1ϵD(u−v)χ0<(u−v)±<ϵϕ. | (3.13) |
Now, we test (3.6) and (3.7) by η1,ϵ to find that
∫Q+4r(x0,t0)∂t(u−v)η1,ϵdxdt+∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),Dη1,ϵ⟩dxdt=∫Q+4r(x0,t0)η1,ϵdμ. | (3.14) |
One can check that
∂t(u−v)η1,ϵ=±∂t(u−v)min{1,(u−v)±ϵ}=∂t[∫(u−v)±0min{1,sϵ}ds]. |
Since ϕ=0 on K4r×{t0} and u=v on K4r×{t0−r2}, integration by parts gives
∫Q+4r(x0,t0)∂t(u−v)η1,ϵdxdt=∫Q+4r(x0,t0)∂t[∫(u−v)±0min{1,sϵ}ds]ϕdxdt=∫Q+4r(x0,t0)∫(u−v)±0min{1,sϵ}ds ∂t{−ϕ(t)}dxdt. | (3.15) |
We obtain from (3.13) and (3.14) that
∫Q+4r(x0,t0)∫(u−v)±0min{1,sϵ}ds ∂t{−ϕ(t)}dxdt+1ϵ∫Q+4r(x0,t0)⟨a(Du,x)−a(Dv,x),D(u−v)χ0<(u−v)±<ϵϕ⟩dxdt=∫Q+4r(x0,t0)η1,ϵdμ. | (3.16) |
By Lebesgue dominated convergence theorem, we get
∫Q+4r(x0,t0)∫(u−v)±0min{1,sϵ}ds ∂t{−ϕ(t)}dxdtϵ→0→∫Q+4r(x0,t0)(u−v)± ∂t{−ϕ(t)}dxdt. |
On the other hand, by ellipticity condition (1.2), we have that
0≤1ϵ∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)χ0<(u−v)±<ϵϕ⟩dxdt. |
Since 0≤η1,ϵ≤1, we find that
∫Q+4r(x0,t0)η1,ϵdμ≤|μ|(Q+4r(x0,t0)). |
So we obtain from (3.16) that
∫Q+4r(x0,t0)(u−v)± ∂t{−ϕ(t)}dxdt≤|μ|(Q+4r(x0,t0)). |
By letting m→∞ for ϕ, we find that
∫K4r(x0)(u(x,τ)−v(x,τ))±dx≤|μ|(Q+4r(x0,t0)). |
Since τ∈(t0−r2,t0) was chosen arbitrarily, this proves the claim (3.11).
Since ϕ is non-increasing, we have that ∂t{−ϕ(t)}≥0, which implies that
∫Q+4r(x0,t0)∫(u−v)±0min{1,sϵ}ds ∂t{−ϕ(t)}dxdt≥0(ϵ>0). |
So it follows from (3.16) that
1ϵ∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)χ0<(u−v)±<ϵϕ⟩dxdt≤∫Q+4r(x0,t0)η1,ϵdμ. |
Since 0≤η1,ϵ≤1, we also obtain from (3.13) that
∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),Dη1,ϵ⟩dxdt≤|μ|(Q+4r(x0,t0)). | (3.17) |
We next claim that
∫Q+4r(x0,t0)|Du−Dv|2(β+|u−v|)ν≤cβ1−νν−1|μ|(Q+4r(x0,t0)) | (3.18) |
for β>0 and ν>1. To this end, for ϵ>0, we test (3.6) and (3.7) by
η2,ϵ=η1,ϵ(β+(u−v)±)ν−1, | (3.19) |
which implies that
∫Q+4r(x0,t0)∂t(u−v)η2,ϵdxdt+∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),Dη2,ϵ⟩dxdt=∫Q+4r(x0,t0)η2,ϵdμ. | (3.20) |
By the same reasoning for (3.15), we get
∫Q+4r(x0,t0)∂t(u−v)η2,ϵdxdt=∫Q+4r(x0,t0)∫(u−v)±0min{1,s/ϵ}(β+s)ν−1ds ∂t{−ϕ(t)}dxdt. |
Since ϕ is non-increasing, we have that ∂t{−ϕ(t)}≥0. So the above equality and (3.11) give that
supϵ>0∫Q+4r(x0,t0)∂t(u−v)η2,ϵdxdt≤β1−νsupτ∈(t0−r2,t0)∫K4r(x0)|u(x,τ)−v(x,τ)|dx≤β1−ν|μ|(Q+4r(x0,t0)). |
One can compute that
∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),Dη2,ϵ⟩dxdt=∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),Dη1,ϵ⟩1(β+(u−v)±)ν−1dxdt+(1−ν)∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)±⟩η1,ϵ(β+(u−v)±)νdxdt. |
Here, we have from (3.17) that
∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),Dη1,ϵ⟩1(β+(u−v)±)ν−1dxdt≤β1−ν∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),Dη1,ϵ⟩dxdt≤β1−ν|μ|(Q+4r(x0,t0)) |
and
|∫Q+4r(x0,t0)η2,ϵdμ|≤β1−ν|μ|(Q+4r(x0,t0)). |
With the above four estimates, we find from (3.20) that
(ν−1)∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)±⟩(β+(u−v)±)ν−1η1,ϵdxdt≤3β1−ν|μ|(Q+4r(x0,t0)). | (3.21) |
By the definition of η1,ϵ, one can see that
∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)±⟩(β+(u−v)±)ν−1η1,ϵdxdt=∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)±⟩(β+(u−v)±)ν−1±min{1,(u−v)±ϵ}ϕdxdt=∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)⟩(β+|u−v|)ν−1min{1,(u−v)±ϵ}ϕdxdt, |
which implies that
∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)±⟩(β+(u−v)±)ν−1η1,ϵdxdtϵ→0→∫Q+4r(x0,t0)⟨a(Du,x,t)−a(Dv,x,t),D(u−v)⟩(β+|u−v|)ν−1ϕdxdt. |
By letting τ→t0 and m→∞, the claim (3.18) follows from (3.12) and (3.21).
Choose β=(−∫Q+4r(x0,t0)|u−v|n+1ndxdt)nn+1 and ν=n+1n. Then by the paraoblic Sobolev embedding (see for instance [14, Chapter 1, Proposition 3.1]), we get
β≤c(n,q)[−∫Q+4r(x0,t0)|Du−Dv|dxdt(supτ∈(t0−r2,t0)∫K4r(x0)|u(x,τ)−v(x,τ)|dx)1n]nn+1. |
It follows from (3.11) that
β≤c[|μ|(Q+4r(x0,t0))]1n+1(−∫Q+4r(x0,t0)|Du−Dv|dxdt)nn+1. | (3.22) |
By Hölder's inequaltiy, (3.17) and (3.22), we obtain that
−∫Q+4r(x0,t0)|Du−Dv|dxdt=−∫Q+4r(x0,t0)|Du−Dv|(β+|u−v|)ν2(β+|u−v|)ν2dxdt≤[−∫Q+4r(x0,t0)|Du−Dv|2(β+|u−v|)νdxdt]12[−∫Q+4r(x0,t0)(β+|u−v|)νdxdt]12≤c(|μ|(Q+4r(x0,t0))|Q+4r(x0,t0)|β1−ν)12βν2≤c[{|μ|(Q+4r(x0,t0))}n+2n+1|Q+4r(x0,t0)|(−∫Q+4r(x0,t0)|Du−Dv|dxdt)nn+1]12. |
Since |Q+4r(x0,t0)|≥crn+2, the lemma follows.
We also prove the comparison estimate between Dv and Dw as follows.
Lemma 3.3. Under the assumption (3.5), we have that
−∫Q+4r(x0,t0)|Dv−Dw|2dxdt≤c[(oscT4r(x′0,t0)Dx′ψ)2+r2−∫Q+4r(x0,t0)|∂tψ|2dxdt]. |
Proof. With v−w+ψ−Dx′ψ(x′0,t0)⋅x′, test (3.7) and (3.8). Fix τ∈(t0−16r2,t0). Then
0=∫τt0−16r2∫K+4r(x0)∂t(v−w)(v−w+ψ−Dx′ψ(x′0,t0)⋅x′)dxdt+∫τt0−16r2∫K+4r(x0)⟨a(Dv,x1,x′,t)−a(Dw,x1,x′,t),Dv−Dw+Dxψ−Dxψ(x′0,t0)⟩dxdt. |
Recall that ψ(x1,x′,t)=ψ(x′,t). It follows from (1.2) and Young's inequality that
∫τt0−16r2∫K+4r(x0)⟨a(Dv,x1,x′,t)−a(Dw,x1,x′,t),Dv−Dw+Dxψ−Dxψ(x′0,t0)⟩dxdt≥λ2∫τt0−16r2∫K+4r(x0)|Dv−Dw|2dxdt−c∫τt0−16r2∫K+4r(x0)|Dxψ−Dxψ(x′0,t0)|2dxdt≥λ2∫τt0−16r2∫K+4r(x0)|Dv−Dw|2dxdt−c{τ−(t0−16r2)}|K+4r(x0)|(oscT4r(x′0,t0)Dx′ψ)2. |
By a direct calculation,
∫τt0−16r2∫K+4r(x0)∂t(v−w)(v−w+ψ−Dx′ψ(x′0,t0)⋅x′)dxdt=∫τt0−16r2∫K+4r(x0)∂t{(v−w+ψ−Dx′ψ(x′0,t0)⋅x′)22}dxdt−∫τt0−16r2∫K+4r(x0)∂tψ{v−w+ψ−Dx′ψ(x′0,t0)⋅x′}dxdt. |
From Young's inequality, we get that
∫τt0−16r2∫K+4r(x0)∂t(v−w)(v−w+ψ−Dx′ψ(x′0,t0)⋅x′)dxdt≥∫K+4r(x0){v(x,τ)−w(x,τ)+ψ(x,τ)−Dx′ψ(x′0,t0)⋅x′}22dx−∫τt0−16r2∫K+4r(x0){v(x,t)−w(x,t)+ψ(x,t)−Dx′ψ(x′0,t0)⋅x′}24{τ−(t0−16r2)}dxdt−c{τ−(t0−16r2)}∫τt0−16r2∫K+4r(x0)|∂tψ|2dxdt. |
Thus we find that
λ2∫τt0−16r2∫K+4r(x0)|Dv−Dw|2dxdt+∫K+4r(x0){v(x,τ)−w(x,τ)+ψ(x,τ)−Dx′ψ(x′0,t0)⋅x′}22dx≤∫τt0−16r2∫K+4r(x0){v(x,t)−w(x,t)+ψ(x,t)−Dx′ψ(x′0,t0)⋅x′}24{τ−(t0−16r2)}dxdt+c{τ−(t0−16r2)}[|K+4r(x0)|(oscT4r(x′0,t0)Dx′ψ)2+∫τt0−16r2∫K+4r(x0)|∂tψ|2dxdt], |
where
∫τt0−16r2∫K+4r(x0){v(x,t)−w(x,t)+ψ(x,t)−Dx′ψ(x′0,t0)⋅x′}24{τ−(t0−16r2)}dxdt≤supτ∈(t0−16r2,t0)∫K+4r(x0){v(x,τ)−w(x,τ)+ψ(x,τ)−Dx′ψ(x′0,t0)⋅x′}24dx. |
Since τ∈(t0−16r2,t0) was arbitrary chosen, we find that
∫Q+4r(x0,t0)|Dv−Dw|2dxdt≤cr2[|K+4r(x0)|(oscT4r(x′0,t0)Dx′ψ)2+∫Q+4r(x0,t0)|∂tψ|2dxdt], |
which proves the lemma.
We use the following reverse Hölder type inequality for comparing Dw and Dg.
Lemma 3.4. Under the assumption (3.5), we have that
(−∫Q+3r(x0,t0)|Dw|2dxdt)12≤c(−∫Q+4r(x0,t0)|Dw|+|Dx′ψ(x′0,t0)|dxdt). |
Proof. We let
γ′=Dx′ψ(x′0,t0)andγ=(0,γ′)=(0,Dx′ψ(x′0,t0)). |
We obtain from (3.5), (3.6) and (3.7) that
v=u=ψonT4r(x′0,t0). | (3.23) |
It follows from (3.10) and (3.23) that
w−γ′⋅x′=v−ψ=0 on T4r(x′0,t0). |
Define ˆw as the zero extension of w−γ′⋅x′ from Q+4r(x0,t0) to Q4r(x0,t0). Then
ˆw={w−γ′⋅x′ in Q+4r(x0,t0),0 in Q4r(x0,t0)∖Q+4r(x0,t0). | (3.24) |
Let 2∗=2nn+2. Then by dividing into two cases (1) Q2ρ(y,s)⊂Rn+1+ and (2) Q2ρ(y,s)⊄Rn+1+, we prove the following assertion that
(−∫Qρ(y,s)|Dˆw|2dxdt)2∗2≤c−∫Q3ρ(y,s)|Dˆw|2∗+|γ|2∗dxdt | (3.25) |
for any Q3ρ(y,s)⊂⊂Q4r(x0,t0).
Choose Q2ρ(y,s)⊂⊂Q4r(x0,t0). First, suppose that Q2ρ(y,s)⊂Rn+1+. Then Q2ρ(y,s)⊂Q+4r(x0,t0). Fix ρ≤r1<r2≤2ρ. Let η∈C∞c(Qr2(y,s)) be a cut-off function with
0≤η≤1,η=1 on Qr1(y,s),|Dη|≤cr2−r1 and |∂tη|≤c(r2−r1)2. | (3.26) |
Fix τ∈(s−r21,s). Take a test function {w−(w)Qr1(y,s)}η2 for (3.8) to find that
0=∫τs−r22∫Kr2(y)∂tw{w−(w)Qr1(y,s)}η2dxdt+∫τs−r22∫Kr2(y)⟨a(Dw,x,t),D[{w−(w)Qr1(y,s)}η2]⟩dxdt. | (3.27) |
One can check that
∫τs−r22∫Kr2(y)∂tw{w−(w)Qr1(y,s)}η2dxdt=∫τs−r22∫Kr2(y)12⋅∂t[{w−(w)Qr1(y,s)}2η2]−{w−(w)Qr1(y,s)}22η∂tηdxdt, |
which implies that
∫Kr2(y)12[{w(x,τ)−(w)Qr1(y,s)}η(x,τ)]2dx≤∫τs−r22∫Kr2(y)[∂tw{w−(w)Qr1(y,s)}η2+c|w−(w)Qr1(y,s)|2|η||∂tη|]dxdt. | (3.28) |
In view of (3.27) and (3.28), we apply the ellipticity condition (1.2) to find that
λ∫τs−r22∫Kr2(y)|Dw|2η2dxdt+∫Kr2(y)12[{w(x,τ)−(w)Qr1(y,s)}η(x,τ)]2dx≤∫τs−r22∫Kr2(y)⟨a(Dw,x,t),η2Dw⟩+∂tw{w−(w)Qr1(y,s)}η2+c|w−(w)Qr1(y,s)|2|η||∂tη|dxdt≤c∫τs−r22∫Kr2(y){|Dw||w−(w)Qr1(y,s)||η||Dη|+|w−(w)Qr1(y,s)|2|η||∂tη|}dxdt. |
First, apply Young's inequality. Then (3.26) gives that
λ2∫τs−r21∫Kr1(y)|Dw|2dxdt+∫Kr1(y,s)12[w(x,τ)−(w)Qr1(y,s)]2dx≤c(r2−r1)2∫τs−r22∫Kr2(y)|w−(w)Qr1(y,s)|2dxdt≤c(r2−r1)2∫Qr2(y,s)|w−(w)Qr2(y,s)|2dxdt, | (3.29) |
where we used that
∫τs−r22∫Kr2(y)|w−(w)Qr1(y,s)|2dxdt≤∫Qr2(y,s)|w−(w)Qr1(y,s)|2dxdt≤c∫Qr2(y,s)|w−(w)Qr2(y,s)|2dxdt, |
which holds from that τ∈(s−r21,s) and ρ≤r1<r2≤2ρ.
Since τ∈(s−r21,s) was arbitrary chosen, we find from (3.28) and (3.29) that
∫Qr1(y,s)|Dw|2dxdt+supτ∈(s−r21,s)∫Kr1(y)[w(x,τ)−(w)Qr1(y,s)]2dx≤c(r2−r1)2∫Qr2(y,s)|w−(w)Qr2(y,s)|2dxdt, |
for any ρ≤r1<r2≤2ρ. By the parabolic Sobolev embedding (see for instance [14, Chapter 1,Proposition 3.1]), we get
∫Qr2(y,s)|w−(w)Qr2(y,s)|2dxdt≤c(∫Qr2(y,s)|Dw|2nn+2dxdt)(supτ∈(s−r22,s)∫Kr2(y)[w(x,τ)−(w)Qr2(y,s)]2dx)2n+2. |
So one can use Young's inequality to find that
∫Qr1(y,s)|Dw|2dxdt+supτ∈(s−r21,s)∫Kr1(y)[w(x,τ)−(w)Qr1(y,s)]2dx≤12supτ∈(s−r22,s)∫Kr2(y)[w(x,τ)−(w)Qr2(y,s)]2dx+c[1(r2−r1)2∫Qr2(y,s)|Dw|2nn+2dxdt]n+2n. |
Let g(θ)=∫Qθ(y,s)|Dw|2dxdt+supτ∈(s−θ,s)∫Kθ(y)[w(x,τ)−(w)Qθ(y,s)]2dx. Then
g(r1)≤12g(r2)+c(r2−r1)2(n+2)n(∫Q2ρ(y,s)|Dw|2nn+2dxdt)n+2n. |
Since ρ≤r1<r2≤2ρ were arbitrary chosen, we obtain from [11, Lemma 4.3] that
g(ρ)≤cρ2(n+2)n(∫Q2ρ(y,s)|Dw|2nn+2dxdt)n+2n, |
which implies that
∫Qρ(y,s)|Dw|2dxdt+supτ∈(s−ρ2,s)∫Kρ(y)[w(x,τ)−(w)Qρ(y,s)]2dx≤cρ2(n+2)n(∫Q2ρ(y,s)|Dw|2nn+2dxdt)n+2n. |
Thus
−∫Qρ(y,s)|Dw|2dx≤c(−∫Q2ρ(y,s)|Dw|2∗dx)22∗. |
Recall from (3.24) that ˆw=w−γ′⋅x′ in Q2ρ(y,s), which implies that
−∫Qρ(y,s)|Dˆw+γ|2dx≤c(−∫Q2ρ(y,s)|Dˆw+γ|2∗dx)22∗. |
So the assertion (3.25) holds for the case Q2ρ(y,s)⊂Rn+1+.
Now, suppose that Q2ρ(y,s)⊄Rn+1+. Then by the fact that Q3ρ(y,s)⊂Q4r(x0,t0),
|Q3ρ(y,s)∩[Q4r(x0,t0)∖Rn+1+]|≥cρn≥c|Q3ρ(y,s)|. | (3.30) |
Fix ρ≤r1<r2≤2ρ. Let η∈C∞c(Qr2(y,s)) be a cut-off function with
0≤η≤1,η=1 on Qr1(y,s),|Dη|≤cr2−r1 and |∂tη|≤c(r2−r1)2. | (3.31) |
Since Q3ρ(y,s)⊂Q4r(x0,t0), it follows from (3.31) that η∈C∞c(Q4r(x0,t0)). We also have from (3.24) that ˆw=0 in Q4r(x0,t0)∖Rn+1+. So we discover that
ˆwη2∈W1,20(Q+4r(x0,t0)). |
Fix τ∈(s−r21,s). Take the test function ˆwη2∈W1,20(Q+4r(x0,t0)) for (3.8) to find that
0=∫τs−r22∫K+r2(y)∂tw[ˆwη2]dxdt+∫τs−r22∫K+r2(y)⟨a(Dw,x,t),D[ˆwη2]⟩dxdt, |
where we used (3.31) and that Q3ρ(y,s)⊂Q4r(x0,t0). By a direct calculation,
∫τs−r22∫K+r2(y)∂tˆw[ˆwη2]dxdt+∫τs−r22∫K+r2(y)⟨a(Dw,x,t)−a(γ,x,t),η2Dˆw⟩dxdt=−∫τs−r22∫K+r2(y){⟨a(Dw,x,t)−a(γ,x,t),ˆw2ηDη⟩+⟨a(γ,x,t),η2Dˆw+ˆw2ηDη⟩}dxdt. |
From (3.24), we have that Dˆw=Dw−γ in Rn+1+. So (1.2) gives that
∫τs−r22∫K+r2(y)∂t{(ˆwη)22}+λ|Dˆw|2η2dxdt≤c∫τs−r22∫K+r2(y){|ˆw|2|η||∂tη|+|η||Dˆw||ˆw||Dη|+|γ|(|η|2|Dˆw|+|ˆw||η||Dη|)}dxdt. |
Since η∈C∞c(Qr2(y,s)), we have from Young's inequality that
∫K+r2(y)|ˆw(x,τ)η(x,τ)|22dx+λ∫τs−r22∫K+r2(y)|Dˆw|2η2dxdt≤c∫Q+r2(y,s){|ˆw|2(|Dη|2+|η||∂tη|)+|γ|2η2}dxdt. | (3.32) |
By (3.24), ˆw=0 in Q4r(x0,t0)∖Rn+1+. Since Q2ρ(y,s)⊂Q4r(x0,t0), we have that
ˆw=0 in Q2ρ(y,s)∖Rn+1+, |
and it follows from (3.32) that
∫Kr2(y)|ˆw(x,τ)η(x,τ)|22dx+∫τs−r22∫Kr2(y)|Dˆw|2η2dxdt≤c∫Qr2(y,s){|ˆw|2(|Dη|2+|η||∂tη|)+|γ|2η2}dxdt. |
Since τ∈(s−r21,s) was arbitrary chosen, we have from (3.31) that
supτ∈(s−r21,s)∫Kr1(y)|ˆw(x,τ)|2dx+∫Qr1(y,s)|Dˆw|2dxdt≤c[1(r2−r1)2∫Qr2(y,s)|ˆw|2dxdt+|γ|2]. | (3.33) |
From (3.30) and that ˆw=0 in Q4r(x0,t0)∖Rn+1+, we have the Sobolev-Poincaré type inequality in [15, Theorem 3.16] to get
−∫Qr2(y,s)|ˆw|2dx=−∫Qr2(y,s)|ˆw(x,τ)|2nn+2|ˆw(x,τ)|4n+2dxdt=−∫ss−r22(−∫Kr2(y)|ˆw(x,τ)|2dx)nn+2(−∫Kr2(y)|ˆw(x,τ)|2dx)2n+2dt≤c(−∫Qr2(y)|Dˆw(x,τ)|2nn+2dx)(supτ∈(s−r22,s)∫Kr2(y)|ˆw(x,τ)|2dx)2n+2. | (3.34) |
So with Young's inequality, one can use the above two inequalities to find that
supτ∈(s−r21,s)∫Kr1(y)|ˆw(x,τ)|2dx+∫Qr1(y,s)|Dˆw|2dxdt≤12supτ∈(s−r22,s)∫Kr2(y)|ˆw(x,τ)|2dx+c[1(r2−r1)2(n+2)n(∫Qr2(y,s)|Dˆw|2nn+2dxdt)n+2n+|γ|2]. |
Let g(θ)=supτ∈(s−θ,s)∫Kθ(y)[ˆw(x,τ)]2dx+∫Qθ(y,s)|Dˆw|2dxdt. Then
g(r1)≤12g(r2)+c[1(r2−r1)2(n+2)n(∫Q3ρ(y,s)|Dˆw|2nn+2dxdt)n+2n+|γ|2]. |
Since ρ≤r1<r2≤2ρ were arbitrary chosen, we obtain from [11, Lemma 4.3] that
g(ρ)≤c[1ρ2(n+2)n(∫Q3ρ(y,s)|Dˆw|2nn+2dxdt)n+2n+|γ|2], |
which implies that
∫Qρ(y,s)|Dˆw|2dxdt+supτ∈(s−ρ2,s)∫Kρ(y)|ˆw(x,τ)|2dx≤c[1ρ2(n+2)n(∫Q3ρ(y,s)|Dˆw|2nn+2dxdt)n+2n+|γ|2]. |
So (3.25) holds for the case Q2ρ(y,s)⊄Rn+1+.
By dividing into two cases (1) Q2ρ(y,s)⊂Rn+1+ and (2) Q2ρ(y,s)⊄Rn+1+, we have the assertion (3.25). Since Q3ρ(y,s)⊂⊂Q4r(x0,t0) in (3.25) was arbitrary chosen, by applying [16, Lemma 3.1] for s=|γ| and χ0=22∗>1 (with a suitable covering argument because the size is 3ρ in the right-hand side of (3.25) not 2ρ), we have that
(−∫Q3r(x0,t0)|Dˆw|2dx)12≤c−∫Q4r(x0,t0)|Dˆw|+|γ|dx. |
Since γ=(0,Dx′ψ(x′0,t0)), the lemma follows from (3.5) and (3.24).
In Lemma 3.4, we obtained the reverse Hölder type inequality. So we can obtain the following comparison estimate for Dw and Dg.
Lemma 3.5. Under the assumption (3.5), we have that
−∫Q+3r(x0,t0)|Dw−Dg|dxdt≤cω(3r)−∫Q+4r(x0,t0)|Dw|+|Dx′ψ(x′0,t0)|dxdt. |
Proof. By using w−g, test (3.8) and (3.9). Then we get
−∫Q+3r(x0,t0)⟨a(Dw,x1,x′0,t0)−a(Dg,x1,x′0,t0),Dw−Dg⟩dxdt≤−∫Q+3r(x0,t0)⟨a(Dw,x1,x′0,t0)−a(Dw,x1,x′,t),Dw−Dg⟩dxdt. |
We obtain from (1.2) and (1.3) that
−∫Q+3r(x0,t0)|Dw−Dg|2dxdt≤cω(3r)−∫Q+3r(x0,t0)|Dw||Dw−Dg|dxdt. |
With Young's inequality and Lemma 3.4, we get that
−∫Q+3r(x0,t0)|Dw−Dg|2dxdt≤c[ω(3r)]2−∫Q+3r(x0,t0)|Dw|2dxdt≤c[ω(3r)]2(−∫Q+4r(x0,t0)|Dw|+|Dx′ψ(x′0,t0)|dxdt)2. |
From Hölder's inequality, the lemma follows.
With Lemma 3.2, Lemma 3.3 and Lemma 3.5, the comparison estimates for Du and Dg will be obtained. We now have the comparison estimate Lemma 3.6 and the excess decay estimate Lemma 2.9, so the remaining proof is similar to the elliptic case in [5]. However, for the sake of the completeness, we give a detailed proof.
Lemma 3.6. Under the assumption (3.5), we have that
−∫Q+3r(x0,t0)|Du−Dg|dxdt≤c[|μ|(Q+4r(x0,t0))rn+1+oscT4r(x′0,t0)Dx′ψ+r2−∫Q+4r(x0,t0)|∂tψ|2dxdt]+cω(4r)−∫Q+4r(x0,t0)|Du|+|Dx′ψ(x′0,t0)|dxdt. |
Proof. By Lemma 3.5,
−∫Q+3r(x0,t0)|Dw−Dg|dxdt≤cω(3r)−∫Q+4r(x0,t0)|Dw|+|Dx′ψ(x′0,t0)|dxdt≤cω(3r)−∫Q+4r(x0,t0)|Du|+|Du−Dw|+|Dx′ψ(x′0,t0)|dxdt. |
In view of Lemma 3.2 and Lemma 3.3,
−∫Q+4r(x0,t0)|Du−Dw|dxdt≤−∫Q+4r(x0,t0)|Du−Dv|+|Dv−Dw|dxdt≤c[|μ|(Q+4r(x0,t0))rn+1+oscT4r(x′0,t0)Dx′ψ+r2−∫Q+4r(x0,t0)|∂tψ|2dxdt]. |
From the above two estimates, the lemma follows.
Recall the definition of G in (3.2), which is an extension of (a1(Dg,x1,x′0,t0),D2g,⋯,Dng). In Lemma 3.6, we obtained an excess decay estimate of G. For an extension U:Q4r(x0,t0)→Rn+1 of (a1(Du,x1,x′0,t0),D2u−D2ψ,⋯,Dnu−Dnψ) which is defined as
U=(u1,⋯,un), | (3.35) |
where
{u1 is the even extension of a1(Du,x1,x′0,t0) from Q+4r(x0,t0) to Q4r(x0,t0),uk(k∈{2,⋯,n}) is the odd extension of Dku−Dkψ from Q+4r(x0,t0) to Q4r(x0,t0), | (3.36) |
we derive an excess decay estimate in the following lemma.
Lemma 3.7. Under the assumption (3.5), we have that
−∫Qρ(x0,t0)|U−(U)Qρ(x0,t0)|dxdt≤c[(ρr)α−∫Q4r(x0,t0)|U−(U)Q4r(x0,t0)|dxdt+ω(4r)(rρ)n−∫Q4r(x0,t0)|U|+|Dx′ψ(x′0,t0)|dxdt]+c(rρ)n[|μ|(Q+4r(x0,t0))rn+1+oscT4r(x′0,t0)Dx′ψ+r2−∫Q+4r(x0,t0)|∂tψ|2dxdt] |
for any .
Proof. We set as
(3.37) |
where
We have from (3.5) and Lemma 2.9 that
(3.38) |
for any . In view of (3.36) and (3.37), we discover that
for any . Since , we have from (3.36) and (3.37) that
(3.39) |
for any . From (3.36) and (3.37), we find that
which implies that
(3.40) |
because in (3.40). For and in (3.35) and (3.37), we have from (3.39) and (3.40) that
On the other hand, Lemma 3.6 gives that
We find from Lemma 3.1 that . So by combining the above two estimates,
(3.41) |
From (3.38) and (3.41), we have that
for any . One can easily check that
and the lemma follows.
The remaining proof is similar to the elliptic case [5], but we give a detailed proof for the completeness. For a weak solution of (1.7), define as
(4.1) |
where
(4.2) |
Lemma 4.1. For any and , we have that
Proof. If , then one can directly check that
and the lemma follows.
We now suppose that . Assume that , which is the interior case. Then . The definition of in (3.2) and (4.1) are different, but the value of in and in from (3.2) and (4.1) differs by at most . Also the value of from (3.2) and (4.1) differs by at most in . So we find from (3.4) that
So the lemma holds when .
On the other hand, if , Lemma 3.7 implies that
where in Lemma 3.7 was defined in (3.35) which is same to that in (4.1). So the lemma holds when .
Now, we prove Theorem 1.1.
Proof of Theorem 1.1. Let be a Lebesgue point of . From Lemma 4.1, we get that
for any . Choose satisfying that
(4.3) |
which implies that
(4.4) |
for any . We choose the constant in (1.6) as
Then from the assumption (1.6) in Theorem 1.1, one can check that
(4.5) |
For , let , , , , , and
(4.6) |
Choose in (4.4). Then we get that
Here, we obtain that
(4.7) |
because of that
Since , one can directly check that
From (4.5), we get
(4.8) |
which implies that for . So from (4.7), we have that
Sum the above inequality over . Then we get
(4.9) |
To simplify the computation, define . Then we have from (4.6) and (4.8) that
We claim that
(4.10) |
To this end, since and , the claim follows holds when . To use induction, we next assume that (4.10) holds when , which means that
(4.11) |
By a direct computation,
which implies that
By using (4.9), we obtain that
We discover from (4.8) and (4.11) that
which implies that
Since , we have that . Also we have that
Thus
and the claim (4.10) holds when . So by an induction, the claim (4.10) holds for .
We have from (4.9) that
which implies that
(4.12) |
By (4.8) and (4.10), we get that
So we find from (4.10) that
In view of (4.12), we get that
By the definition ,
(4.13) |
With the estimate (4.13), it is ready to estimate . Recall from (4.1) and Lemma 3.1 that
(4.14) |
Since , we find from (4.14) that
Since and is an extension defined in (4.2), we have from (4.14) that
By using the above two estimates, we have from (4.13) that
where we used that for . So from the following computation
we obtain that
(4.15) |
From the assumption, is a Lebesgue point of , which implies that
From the definition of , we showed that
So we find that Theorem 1.1 holds.
We explain just a sketch proof for Corollary 1.3 because the proof related to SOLA appears in many papers, say [1, Section 5.2], [17, Section 4.3] and [4, Remark 7]. Suppose that in and in the sense of measures satisfying
for every , where denotes the parabolic closure of . We return to (4.15) in the proof of Theorem 1.1 for the size instead of . For , replace and with and , respectively. Then we find that
for . By sending , we find that
for , which implies that
for , because . So Corollary 1.3 follows, because is a Lebesgue point of .
Ho-Sik Lee : Writing-Original draft preparation; Youchan Kim : Writing-Original draft preparation.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
Y. Kim was supported by the 2023 Research Fund of the University of Seoul. H.-S. Lee was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) through (GRK 2235/2 2021 - 282638148) at Bielefeld University. The authors would like to thank the referee for the careful reading of this manuscript, and offering valuable comments for this manuscript.
The authors declare there is no conflict of interest.
[1] |
F. Duzaar, G. Mingione, Gradient estimates via non-linear potentials, Amer. J. Math., 133 (2011), 1093–1149. https://doi.org/10.1353/ajm.2011.0023 doi: 10.1353/ajm.2011.0023
![]() |
[2] |
T. Kuusi, G. Mingione, Potential estimates and gradient boundedness for nonlinear parabolic systems, Rev. Mat. Iberoam., 28 (2012), 535–576. https://doi.org/10.4171/RMI/684 doi: 10.4171/RMI/684
![]() |
[3] |
T. Kuusi, G. Mingione, The Wolff gradient bound for degenerate parabolic equations, J. Eur. Math. Soc., 16 (2014), 835–892. https://doi.org/10.4171/JEMS/449 doi: 10.4171/JEMS/449
![]() |
[4] |
T. Kuusi, G. Mingione, Riesz potentials and nonlinear parabolic equations, Arch. Ration. Mech. Anal., 212 (2014), 727–780. https://doi.org/10.1007/s00205-013-0695-8 doi: 10.1007/s00205-013-0695-8
![]() |
[5] |
Y. Kim, Y. Youn, Boundary Riesz potential estimates for elliptic equations with measurable nonlinearities, Nonlinear Anal., 194 (2020), 111445. https://doi.org/10.1016/j.na.2019.02.001 doi: 10.1016/j.na.2019.02.001
![]() |
[6] |
Y. Kim, Riesz potential type estimates for nonlinear elliptic equations, J. Differential Equations, 263 (2017), 6844–6884. https://doi.org/10.1016/j.jde.2017.07.031 doi: 10.1016/j.jde.2017.07.031
![]() |
[7] |
S. S. Byun, L. G. Softova, Parabolic obstacle problem with measurable data in generalized Morrey spaces, Z. Anal. Anwend., 35 (2016), 153–171. https://doi.org/10.4171/zaa/1559 doi: 10.4171/zaa/1559
![]() |
[8] |
S. S. Byun, D. Palagachev, L. G. Softova, Global gradient estimates in weighted Lebesgue spaces for parabolic operators, Ann. Acad. Sci. Fenn. Math., 41 (2016), 67–83. https://doi.org/10.5186/aasfm.2016.4102 doi: 10.5186/aasfm.2016.4102
![]() |
[9] |
H. Xu, Existence and blow-up of solutions for finitely degenerate semilinear parabolic equations with singular potentials, Commun. Anal. Mech., 15 (2023), 132–161. https://doi.org/10.3934/cam.2023008 doi: 10.3934/cam.2023008
![]() |
[10] | G. M. Lieberman, Second order parabolic differential equations, World Scientific Publishing Co., Inc., River Edge, NJ, 1996. https://doi.org/10.1142/3302 |
[11] | Q. Han, F. Lin, Elliptic partial differential equations, Second edition, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 2011. |
[12] |
S. S. Byun, Y. Kim, Elliptic equations with measurable nonlinearities in nonsmooth domains, Adv. Math., 288 (2016), 152–200. https://doi.org/10.1016/j.aim.2015.10.015 doi: 10.1016/j.aim.2015.10.015
![]() |
[13] |
S. S. Byun, Y. Kim, Riesz potential estimates for parabolic equations with measurable nonlinearities, Int. Math. Res. Not., 2018 (2018), 6737–6779. https://doi.org/10.1093/imrn/rnx080 doi: 10.1093/imrn/rnx080
![]() |
[14] | E. DiBenedetto, Degenerate parabolic equations, Universitext, Springer-Verlag, New York, 1993. https://doi.org/10.1007/978-1-4612-0895-2 |
[15] | E. Giusti, Direct methods in the calculus of variations, World Scientific Publishing Co., Inc., River Edge, NJ, 2003. https://doi.org/10.1142/9789812795557 |
[16] |
G. Mingione, The Calderón-Zygmund theory for elliptic problems with measure data, Ann. Sc. Norm. Super. Pisa Cl. Sci., 6 (2007), 195–261. https://doi.org/10.2422/2036-2145.2007.2.01 doi: 10.2422/2036-2145.2007.2.01
![]() |
[17] |
T. Kuusi, G. Mingione, New perturbation methods for nonlinear parabolic problems, J. Math. Pures Appl., 98 (2012), 390–427. https://doi.org/10.1016/j.matpur.2012.02.004 doi: 10.1016/j.matpur.2012.02.004
![]() |